Guest-Induced Structural Transformation and Its Effect on the

of Chemistry and Chemical Engineering, Harbin Institute of Technology, Harbin, .... on the principle of DCC to study the effect of surface-confine...
1 downloads 4 Views 1MB Size
Subscriber access provided by Kaohsiung Medical University

C: Surfaces, Interfaces, Porous Materials, and Catalysis

Guest-Induced Structural Transformation and Its Effect on the Redistribution of the Surface Confined Combinatorial Libraries Jiang Sun, and Shengbin Lei J. Phys. Chem. C, Just Accepted Manuscript • DOI: 10.1021/acs.jpcc.8b03215 • Publication Date (Web): 17 May 2018 Downloaded from http://pubs.acs.org on May 17, 2018

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 25 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Guest-Induced Structural Transformation and Its Effect on the Redistribution of the Surface Confined Combinatorial Libraries Jiang Sun a and Shengbin Lei ab* a

School of Chemistry and Chemical Engineering, Harbin Institute of Technology, Harbin,

150080, People's Republic of China, Email: [email protected] b

Tianjin Key Laboratory of Molecular Optoelectronic Science, Department of Chemistry,

School of Science & Collaborative Innovation Center of Chemical Science and Engineering (Tianjin), Tianjin University, Tianjin 300072, People’s Republic of China

ABSTRACT: With the aim of better understanding the dynamic covalent chemistry (DCC) and host-guest interaction at the liquid/solid interface, Schiff base reaction was employed as a medium to investigate the distribution/redistribution of the dynamic covalent library (DCL) toward the presence of surface. In addition, coronene is used as guest molecule to observe the guest-induced structural transformation at the liquid/solid interface. It is found that introduction of hydroxyl group which can form intra and inter-molecular hydrogen bond significantly changes the response of DCL toward both surface and guest molecules at the interface. With the aid of surface introduction of the coronene as guest can significantly amplify specific product.

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

INTRODUCTION The application of “bottom-up” approach, e.g. self-assembly, in the fabrication of a nanopatterned functional surface capable of responding to external stimuli has drawn more attention in recent years due to its diversity and easy accessibility compared to its traditional counterpart, namely “top-down” route1-12. Many an apparatus has been applied to investigate the surface patterning and the principle behind it, among which STM exhibits outstanding advantages owing to its high spatial resolution and mild working environment in comparison with other characterization method13-18. In the realm of construction of supramolecular self-assemblies, host-guest interactions account for a great importance19-22. Generally speaking, host-guest interactions on the surface is mainly classified into two distinct categories, that is, guest accommodation in the host networks23-27 and guest-induced structural transformation of the host matrix28-30. In addition to self-assembly for the formation of supramolecular complexes, on-surface reaction is an emerging alternative for the nanopattern fabrication and synthesis of low dimensional polymers, which has developed, over the past years, to the point where both one-dimensional linear polymers and long-range ordered two-dimensional porous networks, are accurately accessible through elaborate design of the precursor and judicious regulation of the environmental conditions. The commonly seen surface-assisted reactions include, but not limited to, Ullmann coupling, boronic acid condensation, Schiff base reaction, Glaser coupling, etc31-38. Of those, Schiff base reaction is the one extensively studied under atmospheric conditions, whereby a particularly broad palette of 2D porous honeycomb networks with disparate cavity sizes were synthesized. With the development of STM and progress in understanding the mechanism of surface-mediated reaction, rational design and in situ observation of enlightening

ACS Paragon Plus Environment

Page 2 of 25

Page 3 of 25 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

phenomenon at the liquid/solid interface have come into reality. In spite of the status quo that DCC has long since been studied in bulk phase39-42, its investigation at an interface has not been developed until quite recent years43,44. In this respect, Schiff base reaction affords an ideal platform to investigate the surface confined DCC due to the intrinsic reversibility of the C=N linkage under ambient conditions which renders the appealing prospects of “error-checking” and “proof-reading” for the surface-assisted reaction product. For the surface confined DCC, adsorption and assembling of species plays a key role in amplification toward a specific product, and the surface composition could be significantly different from that in the solution, in extreme, if large enough surface area could be provided, it is possible to drive the DCL toward the equilibrium to production of an aimed product. Herein,

2,5-bis(octyloxy)-terephthalaldehyde,

4,4'-biphenyldicarboxaldehyde,

3,3'-dihydroxy-4,4'-biphenyldicarboxaldehyde and 5-aminoisophthalic acid (see Scheme 1) were employed as precursors to investigate the DCC between imine and aldehyde at the octanoic acid/HOPG interface. Besides, the coronene-induced structural transformation was observed for both binary and tri-component DCLs, which contributed to the amplification of specific product in a DCL.

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Scheme 1. (a) Chemical structure of monomers and coronene molecule. (b) possible imine products from the DCLs. RESULTS AND DISCUSSION In this work, binary and tri-component dynamic covalent libraries were established based on the principle of DCC to study the effect of surface-confinement at the octanoic acid/HOPG interface. Besides, guest-induced structural transformation and its effect on the product distribution of a DCL were also investigated. The structure of the reactants and observed products on the surface are shown in Scheme 1a and 1b. 1. Bi-component DCLs at the liquid/solid interface After the mixture of monomer 4,4'-biphenyldicarboxaldehyde (3) and 5-aminoisophthalic acid (1) with a molar ratio of 1:2 was dropcasted onto the freshly-cleaved HOPG surface, STM characterization was carried out, and two discernible close-packed patterns were visualized as shown in Fig. 1. From the large scale STM image in Fig. 1 (a), it is found that both of the obtained assembling patterns, marked as I and II, are constructed by rod-like

ACS Paragon Plus Environment

Page 4 of 25

Page 5 of 25 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

building blocks with different packing mode. To be specific, as is unraveled by the high resolution STM image in Fig. 1 (b), the building unit in pattern I adopts a parallel arrangement, similar to the one in a previous report. 45 The unit cell parameters of pattern I were measured to be a = 2.3 ± 0.1 nm, b = 2.4 ± 0.1 nm, and α = 44 ± 1°. Although the sub-molecular feature is hardly distinguishable, it is plausible to deduce through the unit cell parameters and the assembly fashion that the building unit is the diimine product 7. Based on this, tentative molecular model was proposed and displayed in Fig. 1 (d). From the model, one can see that both of the two carboxylic groups at each end of the diimine product 7 form a pair of hydrogen bonds with the adjacent building unit. Unlike the previous report, 45 pattern I is stabilized exclusively by hydrogen bonding, without solvent involved. As to assembling pattern II, it is a distinct scenario in which every two of the bright features form a dimer, as circled by a red oval shown in Fig. 1 (c). According to the size of the bright features they are asigned to the monoimine product 8. Through attentive observation of the high resolution STM image, it is revealed that the interval between the two building blocks in the dimer is smaller than that between two neighboring dimers. In other words, the interaction between the molecules in the dimer is different from that between the dimers. Therefore, we make an inference that the solvent also plays a role in the stabilization of pattern II, though it is not clearly resolved in the STM image. Thanks to the inequality of the two intervals and the symmetry breaking effect of the surface, chiral assembly domains were observed, the enantiomeric domains are separated by the red lines in Figure 1c. Based on the experimentally measured unit cell parameters (a = 2.0 ± 0.1 nm, b = 3.7 ± 0.2 nm, and α = 87 ± 2°) and the deduced assembling arrangement, tentative molecular model was constructed

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

and presented in Fig. 1 (e). Two monoimine molecules formed a dimer as discussed above through a pair of hydrogen bonds between carboxylic groups, while at the other end the aldahyde group remained intact. Entrapped in the interspace between two adjacent dimers are two co-adsorped n-octanoic acid molecules which contribute to stabilize the adlayer through van der Waals forces.

Figure 1. Large-scale(a) and typical high-resolution STM image of the two different close-packed patterns I

(b) and II (c) formed after the mixture of 3 and 1 (molar ratio 1:2)

was dropcasted onto the HOPG surface. Scale bar = 20 nm.and 5 nm, respectively. Suggested molecular model of structure I (d) and

structure II (e). H-bond linkage was highlighted by

the white rectangle and magnified on the top right corner, respectively. Tunneling conditions: (a) Vbias= 750 mV, Iset= 40 pA; (b) Vbias= 750 mV, Iset= 35 pA; (c) Vbias= 700 mV, Iset= 30 pA.

Different from 3, the addition of hydroxyl groups on the backbone of 4 is expected to cause variation in intermolecular interactions, thus changes the assembling behavior of the products. After the mixture of monomer 4 and 1 with a molar ratio of 1:2 was dropcasted onto the

ACS Paragon Plus Environment

Page 6 of 25

Page 7 of 25 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

freshly-cleaved HOPG surface, the sample was heated at 60 °C for 30 minutes. STM characterization revealed a zigzag motif as shown in Fig. 2. From the large-scale STM image in Fig. 2 (a), it is found that the surface was covered exclusively by the zigzag assembly of the product. This is, to a large extent, an indication that only one species (mono- or diimine) was generated at the interface and form the aforementioned zigzag pattern. This is verified by the high resolution STM image in Fig. 2 (b), in which the details of the zigzag pattern are unambiguously illustrated. The zigzag structure consisted of rod-like building blocks which are interconnected in an orthogonal way. Based on the unit cell parameters which are measured to be a = 1.9 ± 0.1 nm, b = 2.8 ± 0.1 nm, and α = 86 ± 2° and the arrangement details, a tentative molecular model is proposed and presented in Fig. 2 (c). From the model, it is revealed that the product generated here was diimine 9 and the diimines were interlinked by hydrogen bonds. Besides, unlike 3 and 1 which can react and lead to ordered assembly on the surface at room temperature, Schiff base reaction between 4 and 1 cannot complete on the same condition. Room temperature reaction normally leads to disordered assembly of mixtures of moniimine 10 and diimine 9, and this is attributed to the stronger intermolecular hydrogen bonds between O−H⋅⋅⋅O between hydroxyl groups in 4, and hydroxyl group and carboxyl group in moniimine 10 and diimine 9, which leads to lower mobility on the surface.

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Figure 2. Large-scale (a) and typical high-resolution (b) STM image of the zigzag pattern. Scale bar = 20 nmand 5 nm, respectively. (c) Proposed molecular model of the self-assembling monolayer. An enlarged model was inserted in the top right corner. Tunneling conditions: (a) Vbias= 600 mV, Iset= 35 pA; (b) Vbias= 650 mV, Iset= 30 pA. 2. Tri-component DCLs at the liquid/solid interface After the mixture of monomer 3, 2 and 1 with a molar ratio of 1:1:2 was dropcasted onto the freshly-cleaved HOPG surface, STM characterization was carried out, and a close-packed structure was observed as shown in Fig. 3. From the discussion above and previous report45, we know that both 3 and 2 can react with 1 at the octanoic acid/HOPG interface and the products self-assemble on the surface. Here, large-scale STM image shown in Fig. 3 (a) reveals exclusively the assembly pattern of diimine 5 at the interface. With the aid of high resolution STM image in Fig. 3 (b), in which the octyloxy group is clearly discernible, this attribution is further confirmed. Extending the reaction time at the interface to 12 hours disclosed no prominent distinction as shown in Fig. 3 (c), indicating that the preformed adlayer is thermodynamically stable. Nonetheless, there is a subtle nuance worth noticing, that is, the interval between neighboring lamellae is widened at some area as pointed by the white arrows in Fig. 3 (c). The abundance of such widened intervals increases when the reaction time was prolonged to 24 hours, as

ACS Paragon Plus Environment

Page 8 of 25

Page 9 of 25 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

shown in Fig. 3 (d). Careful inspection of the high resolution STM image in Fig. 3 (d) discloses that some longer rods (circled by red ovals) are embedded in the assembly of product 5, giving rise to the widened interval between two adjacent lines, and the longer rods are justifiably attributed to the diimine product 7 formed by 3 and 1 based on the way they interact with adjacent building blocks.

Figure 3. Large-scale (a) and typical high-resolution (b) STM image of the close-packed pattern after the mixture of 3, 2 and 1 (molar ratio 1:1:2) was dropcasted onto the HOPG surface. Scale bar = 20 nm and 5 nm, respectively. (c) Large-scale STM image 12h after the deposition of the monomer mixture. Some widened rows are pointed by the white arrows. Scale bar = 20 nm. (d) High-resolution STM image 24h after the deposition of the monomer mixture. Scale bar = 5 nm. The longer rods embedded in the lamellae are highlighted with the red oval. Tunneling conditions: (a) Vbias= 750 mV, Iset= 38 pA; (b) Vbias= 650 mV, Iset= 35 pA; (c) Vbias= 700 mV, Iset= 40 pA; (d) Vbias= 600 mV, Iset= 38 pA.

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

As in the case of a tri-component DCL containing 4, 2 and 1 with a molar ratio of 1:1:2, STM characterization reveals also exclusively the assembly of diimine product 5 after the sample was heated at 60°C for 30 minutes. As time passed by, no other species emerged, suggesting that the surface DCL had reached equilibrium. The contact with a surface of both tri-component DCLs leads to the amplification of product 5 over other possible products (6, 9 and 10), possibly due to the preferential adsorption and assembly of this specific product at the HOPG/octanoic acid interface.

Figure 4. (a) Large-scale STM image of the close-packed pattern of product 5. Scale bar = 20 nm. (b) Typical high-resolution STM image in which the octyloxy group of product 5 is discernible. Scale bar = 5 nm. Tunneling conditions: (a) Vbias= 700 mV, Iset= 38 pA; (b) Vbias= 650 mV, Iset= 38 pA.

3. Metathesis at the octanoic acid/HOPG interface The mixture of monomer 3 and 1 with a molar ratio of 1:2 was dropcasted onto the freshly-cleaved HOPG surface, then 2 was deposited onto the surface in a amount ensuring the consequent molar ratio of the three monomers at the interface was 1:2:1. STM characterization visualized a close-packed pattern as shown in Fig. 5. From the large scale STM image in Fig. 5 (a), it is unraveled that the preformed adlayer (both diimine 7 and

ACS Paragon Plus Environment

Page 10 of 25

Page 11 of 25 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

monoimine 8 assembly domains) totally disappeared from the surface and was replaced by a close-packed assembly structure which was ascribed to the diimine product 5 based on the arrangement fashion and the unit cell parameters. However, it is worth noticing that some individual product 7 formed by 3 and 1 were embedded in the monolayer, giving rise to the widened separation between two neighboring lamellae at some area as presented in Fig. 5 (b). Compared to the transimination with amino compound, the phenomenon observed here can be explained by the metathesis between imine and aldehyde compound at the interface.

Figure 5. (a) Large-scale STM image of the close-packed pattern. Scale bar = 20 nm. (b) High-resolution STM image showing the longer rod (circled by oval) embedded in the original structure. Scale bar = 5 nm. Tunneling conditions: (a) Vbias= 720 mV, Iset= 38 pA; (b) Vbias= 700 mV, Iset= 30 pA.

After the formation of the zig-zag pattern of product 9, various amount of 2 was deposited onto the surface. STM characterization reveals no change of the zigzag motif as shown in Fig. 6, indicating that metathesis did not occur at the interface. In contrast to the co-deposition of the three monomers where the surface was covered exclusively by the diimine product 5, it is revealed here that once the zigzag structure of 9 preformed at the interface, the introduction of

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

2 cannot induce the alteration of the surface assembly pattern, that is, metathesis does not take place under such circumstance. This may be caused by the synergy effect of the intramolecular hydrogen bond formed between OH--O in 4 moiety and the high adsorption energy, which enhances the stability of the zigzag motif on the surface.

Figure 6. Large-scale STM image after 2 was dropcasted onto the surface predeposited with 4 and 1 with the molar ratio of (a) 5:1:2 and (b) 10:1:2. Scale bar = 20 nm for (a) and 15 nm for (b). In both cases, only diimine 9 formed by 4 and 1 was observed. Tunneling conditions: (a) Vbias= 680 mV, Iset= 30 pA; (b) Vbias= 700 mV, Iset= 35 pA.

3. Guest promoted product selection from DCLs For the study of guest promoted amplification toward specific products, the mixture of monomer 3 and 1 with a molar ratio of 1:2 was first dropcasted onto the freshly-cleaved HOPG surface, after formation of the assemblies of product 7 and 8, a droplet of coronene (COR) was deposited. STM characterization reveals the COR-induced structural transformation as shown in Fig. 7 soon after the deposition. The large scale STM image in Fig. 7 (a) unravels that the surface was covered predominantly by the COR-induced Kagome structure of product 7, and the preformed assembling arrangement of pattern I and II vanished, suggesting that the Kagome structure was more stable than both pattern I and II. It is well

ACS Paragon Plus Environment

Page 12 of 25

Page 13 of 25 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

known that the strong interaction between COR with the cyclic pores formed by isophthalate moieties leads to formation of stable host-guest architecture, and also can induce structural transformation of surface assemblies.[46] Thus the amplification of product 7 by the addition of COR can also be attributed to the strong interaction between COR and isophthalate moieties of product 7. In addition, as shown in Fig. 7 (b), fuzzy features can be detected in the triangular cavity enclosed by the diimine product 7, which are attributed to COR or other small molecule entrapped in the pores. The unit cell parameters are measured to be a = 4.4 ± 0.1 nm, b = 4.4 ± 0.1 nm, and α = 60 ± 1°. Besides, in a control experiment, after the COR-induced Kagome structure was acquired, dialdehyde 2 was dropcasted onto the surface, however, the surface assembly structure was not altered, further confirming the stability of the Kagome motif, which prohibits the imine metathesis.

Figure 7. (a) Large-scale STM image of the Kagome structure of product 7 formed after addition of COR. Scale bar = 20 nm. (b) Typical high-resolution STM image with the unit cell superimposed. Scale bar = 5 nm. (c) Suggested molecular model of the bicomponent Kagome structure. Tunneling conditions: (a) Vbias= 800 mV, Iset= 30 pA; (b) Vbias= 850 mV, Iset= 28 pA.

Similar guest induced structural transformation is also observed on the assembly of 9. After the mixture of monomer 4 and 1 with a molar ratio of 1:2 was dropcasted onto the freshly-cleaved HOPG surface, the sample was heated at 60°C for 30 minutes. Then COR was

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

deposited onto the surface. COR-induced Kagome pattern was observed and the parental zigzag motif disappeared as presented in Fig. 8 (a), suggesting that the Kagome structure was more stable than the zigzag pattern. Similarly, the triangular pore was occupied by fuzzy feature attributed to the COR molecule or other small organic species at the interface. The unit cell parameters are measured to be a = 4.4 ± 0.1 nm, b = 4.4 ± 0.1 nm, and α = 60 ± 1°. Notably, some individual COR molecules entrapped in the triangular cavity were clearly resolved due to the existence of -OH group on the diimine product 9, as circled by red ellipse in Fig. 8 (b), which contributed to the immobilization of the COR molecule.

Figure 8. Large-scale (a) and typical high-resolution (b) STM image of the Kagome structure. Scale bar = 20 nm and 5 nm, respectively. (c) Suggested molecular model of the Kagome structure. Tunneling conditions: (a) Vbias= 850 mV, Iset= 30 pA; (b) Vbias= 900 mV, Iset= 25 pA.

Efforts were made for the extended study of guest promoted product amplification from a tri-component DCL. To this end, a 2µL drop of mixture solution of 1, 2, and 3 with a molar ratio of 2:1:1 was applied on the HOPG surface, and STM characterization unraveled the assembling structure of product 5 as shown in Fig. 9 (a). Afterwards, a 2µL drop of saturated COR solution was introduced on top of the above adlayer, and guest-induced structural transformation from close-packed pattern to truss and Kagome pattern was visualized. Through careful observation of Fig. 9 (b), the octyloxy group was discernible, confirming that

ACS Paragon Plus Environment

Page 14 of 25

Page 15 of 25 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

the guest-induced structure was composited by the coadsorption of COR with product 5, not with product 7.

Figure 9. (a) Large-scale STM image of the close-packed pattern. Scale bar = 20 nm. (b) COR-induced structural transformation. Scale bar = 10 nm. Tunneling conditions: (a) Vbias= 750 mV, Iset= 38 pA; (b) Vbias= 850 mV, Iset= 25 pA.

In comparison, when the saturated solution of COR is premixed (1:1 vol:vol) with the tri-component mixture solution of 1, 2, and 3 with a molar ratio of 2:1:1, it was unveiled that the surface was covered by a Kagome structure formed by the coadsorption of COR with product 7 (Fig. 10a). The triangular cavity of this Kagome structure was occupied by fuzzy features, and the unit cell parameters are measured to be a = 4.4 ± 0.1 nm, b = 4.4 ± 0.1 nm, and α = 60 ± 1°. After 24 hours under ambient conditions, the surface was still exclusively covered by the Kagome pattern formed by COR and product 7 as shown in Fig. 10 (b), indicating the stability of this Kagome structure.

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Figure 10. (a) Kagome structure formed after the mixture solution was dropcasted onto the HOPG surface. Scale bar = 5 nm. (b) The surface was still exclusively covered by the Kagome pattern formed by COR and product 7 after 24 hours under ambient conditions. Scale bar = 5 nm. Tunneling conditions: (a) Vbias= 900 mV, Iset= 30 pA; (b) Vbias= 900 mV, Iset= 25 pA.

As to the study of guest promoted product selection from a tri-component DCL containing 1, 2 and 4, the protocol was similar as above-mentioned. The mixture solution of 1, 2, and 4 with a molar ratio of 2:1:1 was mixed with the saturated solution of COR (1:1 vol:vol), and then a 2µL drop of the mixture solution was deposited onto the HOPG surface which was preheated to 60 °C. The sample was further annealed at 60 °C for 30 min in an oven, and then characterized by STM. Through the typical STM image as displayed in Fig. 11 (a), it was unveiled that the surface was covered by the Kagome structure, while the triangular cavity of the Kagome structure was unoccupied, indicating that the Kagome motif was formed by the coadsorption of COR with product 5, not with product 9. Extending the annealing duration to 2 hours, the surface was still exclusively covered by the Kagome pattern formed by COR and product 5 as illustrated in Fig. 11 (b), suggesting the stability of this Kagome structure in this case.

ACS Paragon Plus Environment

Page 16 of 25

Page 17 of 25 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Figure 11. (a) Kagome structure formed after the mixture solution was dropcasted onto the HOPG surface. Scale bar = 5 nm. (b) The surface was still exclusively covered by the Kagome pattern formed by COR and product 5 when the annealing duration was extended to 2 hours. Scale bar = 5 nm. Tunneling conditions: (a) Vbias= 900 mV, Iset= 28 pA; (b) Vbias= 900 mV, Iset= 28 pA.

CONCLUSIONS In summary, the investigation of DCC between imine and aldehyde was carried out at the octanoic acid/HOPG interface using dialdehydes 2, 3, 4 and amine 1 as precursors. Meanwhile, COR was used as guest molecule to investigate the guest-promoted structural transformation and its effect on the redistribution of the combinatorial libraries. It was found that DCC took place between 2 and the diimine 7, while the zigzag pattern of the diimine 9 was not influenced by the addition of 2. Besides, the COR-induced structural transformation was observed for both binary and tri-component DCLs, contributing to the amplification of specific product in a DCL. The result in this work broadens the range of DCC at the liquid/solid interface and sheds light on the guest-induced structural transformation and its effect on the redistribution of the combinatorial libraries.

MATERIALS AND METHODS

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 18 of 25

Coronene, n-octanoic acid, 2,5-bis(octyloxy)-terephthalaldehyde, 5-aminoisophthalic acid, 4,4'-biphenyldicarboxaldehyde,

and

3,3'-dihydroxy-4,4'-biphenyldicarboxaldehyde

were

purchased from J&K and used without further treatment. For the guest molecule induced structural transition, coronene was first dissolved in octanoic acid at room temperature (about 25°C), and used either directly as a saturated solution, or diluted accordingly. The sample was prepared by casting a droplet of solution containing coronene molecule onto the HOPG surface on which the monomers were predeposited and the imine products were visualized by STM. The concentration of monomer solution is 0.1 mg/mL, unless otherwise specified. The STM observations were carried out under ambient conditions at the liquid/solid interface for all cases by using an Agilent 5100 Scanning Probe Microscopy, and the tips were mechanically cut Pt/Ir (80/20) wires. All the STM images were recorded using constant current mode, and the specific tunneling conditions were illustrated in the corresponding figure captions. The calibration of the STM images was performed by using HOPG lattice with atomic resolution. ACKNOWLEDGMENT This work was supported by the National Science Foundation of China (21572157).

REFERENCES

(1) Nath, K. G.; Ivasenko, O.; MacLeod, J. M.; Miwa, J. A.; Wuest, J. D.; Nanci, A.; Perepichka, D. F.; Rosei, F. Crystal Engineering in Two Dimensions: An Approach to Molecular Nanopatterning. J. Phys. Chem. C 2007, 111, 16996-17007. (2) Ahn, S.; Matzger, A. J. Six Different Assemblies from One Building Block:

ACS Paragon Plus Environment

Page 19 of 25 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Two-Dimensional Crystallization of an Amide Amphiphile. J. Am. Chem. Soc. 2010, 132, 11364-11371. (3) Barth, J. V.; Costantini, G.; Kern, K. Engineering Atomic and Molecular Nanostructures at Surfaces. Nature 2005, 437, 671-679. (4) Mali, K. S.; Wu, D.; Feng, X.; Müllen, K.; Van der Auweraer, M.; De Feyter, S. Scanning Tunneling Microscopy-Induced Reversible Phase Transformation in the Two-Dimensional Crystal of a Positively Charged Discotic Polycyclic Aromatic Hydrocarbon. J. Am. Chem. Soc. 2011, 133, 5686-5688. (5) Cometto, F. P.; Kern, K.; Lingenfelder, M. Local Conformational Switching of Supramolecular Networks at the Solid/Liquid Interface. ACS Nano 2015, 9, 5544-5550. (6) Lei, S. B.; Deng, K.; Yang, Y. L.; Zeng, Q. D.; Wang, C.; Jiang, J. Z. Electric Driven Molecular Switching of Asymmetric Tris(phthalocyaninato) Lutetium Triple-Decker Complex at the Liquid/Solid Interface. Nano Lett. 2008, 8, 1836-1843. (7) Kudernac, T.; Ruangsupapichat, N.; Parschau, M.; Macia, B.; Katsonis, N.; Harutyunyan, S. R.; Ernst, K. H.; Feringa, B. L. Electrically Driven Directional Motion of a Four-Wheeled Molecule on a Metal Surface. Nature 2011, 479, 208-211. (8) Pawin, G.; Wong, K. L.; Kwon, K. Y.; Bartels, L. A Homomolecular Porous Network at a Cu(111) Surface. Science 2006, 313, 961-962. (9) Uemura, S.; Aono, M.; Komatsu, T.; Kunitake, M. Two-Dimensional Self-Assembled Structures of Melamine and Melem at the Aqueous Solution-Au(111) Interface. Langmuir 2011, 27, 1336-1340. (10) Hieulle, J.; Silly, F. Localized Intermolecular Electronic Coupling in Two-Dimensional

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 20 of 25

Self-Assembled 3,4,9,10-perylenetetracarboxylic Diimide Nanoarchitectures. J. Mater. Chem. C 2013, 1, 4536-4539. (11) Tait, S. L. Function Follows Form: Exploring Two-Dimensional Supramolecular Assembly at Surfaces. ACS Nano 2008, 2, 617-621. (12) Ruben, M.; Payer, D.; Landa, A.; Comisso, A.; Gattinoni, C.; Lin, N.; Collin, J. P.; Sauvage,

J.

P.;

De

Vita,

A.;

Kern,

K.

2D

Supramolecular

Assemblies

of

Benzene-1,3,5-triyl-tribenzoic Acid: Temperature-induced Phase Transformations and Hierarchical Organization with Macrocyclic Molecules. J. Am. Chem. Soc. 2006, 128, 15644-15651. (13) Gutzler, R.; Sirtl, T.; Dienstmaier, J. r. F.; Mahata, K.; Heckl, W. M.; Schmittel, M.; Lackinger, M. Reversible Phase Transitions in Self-Assembled Monolayers at the Liquid– Solid Interface: Temperature-Controlled Opening and Closing of Nanopores. J. Am. Chem. Soc. 2010, 132, 5084-5090. (14) Yau, S. L.; Kim, Y. G.; Itaya, K. High-Resolution Imaging of Aromatic Molecules Adsorbed on Rh(111) and Pt(111) in Hydrofluoric Acid Solution: In Situ STM Study. J. Phys. Chem. B 1997, 101, 3547-3553. (15) Yamada, T.; Shibuta, M.; Ami, Y.; Takano, Y.; Nonaka, A.; Miyakubo, K.; Munakata, T. Novel Growth of Naphthalene Overlayer on Cu(111) Studied by STM, LEED, and 2PPE. J. Phys. Chem. C 2010, 114, 13334-13339. (16) Bartels, L. Tailoring Molecular Layers at Metal Surfaces. Nat. Chem. 2010, 2, 87-95. (17) Matsunaga, S.; Yamada, T.; Kobayashi, T.; Kawai, M. Scanning Tunneling Microscope Observation of the Phosphatidylserine Domains in the Phosphatidylcholine Monolayer.

ACS Paragon Plus Environment

Page 21 of 25 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Langmuir 2015, 31, 5449-5455. (18) Cyr, D. M.; Venkataraman, B.; Flynn, G. W. STM Investigations of Organic Molecules Physisorbed at the Liquid-Solid Interface. Chem. Mater. 1996, 8, 1600-1615. (19) Iritani, K.; Tahara, K.; De Feyter, S.; Tobe, Y. Host-Guest Chemistry in Integrated Porous Space Formed by Molecular Self-Assembly at Liquid-Solid Interfaces. Langmuir 2017, 33, 4601-4618. (20) Teyssandier, J.; Feyter, S. D.; Mali, K. S. Host-guest chemistry in two-dimensional supramolecular networks. Chem. Commun. 2016, 52, 11465-11487. (21) Spillmann, H.; Kiebele, A.; Stöhr, M.; Jung, T.; Bonifazi, D.; Cheng, F.; Diederich, F. A Two-Dimensional Porphyrin-Based Porous Network Featuring Communicating Cavities for the Templated Complexation of Fullerenes. Adv. Mater. 2006, 18, 275-279. (22) Zhang, X.; Chen, T.; Yan, H. J.; Wang, D.; Fan, Q. H.; Wan, L. J.; Ghosh, K.; Yang, H. B.; Stang, P. J. Engineering of Linear Molecular structures by a Hydrogen-Bond-Mediated Modular and Flexible Host-Guest Assembly. ACS Nano 2010, 4, 5685-5692. (23) Lee, S. L.; Lin, C. H.; Cheng, K. Y.; Chen, Y. C.; Chen, C. H. Stability of Guest-Incorporated 2D Molecular Networks. J. Phys. Chem. C 2016, 120, 25505-25510. (24) Sun, J.; Zhou, X.; Lei, S. B. Host-Guest Architectures with a Surface Confined Imine Covalent Organic Framework as Two-Dimensional Host Networks. Chem. Commun. 2016, 52, 8691-8694. (25) Schmaltz, B.; Rouhanipour, A.; Räder, H. J.; Pisula, W.; Müllen, K. Filling the Cavity of Conjugated Carbazole Macrocycles with Graphene Molecules: Monolayers Formed by Physisorption Serve as a Surface for Pulsed Laser Deposition. Angew. Chem., Int. Ed. 2009,

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

48, 720-724. (26) Meier, C.; Landfester, K.; Künzel, D.; Markert, T.; Groß, A.; Ziener, U. Hierarchically Self-Assembled Host-Guest Network at the Solid-Liquid Interface for Single-Molecule Manipulation. Angew. Chem., Int. Ed. 2008, 47, 3821-3825. (27) Schull, G.; Douillard, L.; Fiorini-Debuisschert, C. l.; Charra, F.; Mathevet, F.; Kreher, D.; Attias, A.-J. Single-Molecule Dynamics in a Self-Assembled 2D Molecular Sieve. Nano Lett. 2006, 6, 1360-1363. (28) Blunt, M.; Lin, X.; Gimenez-Lopez, M.; Schröder, M.; Champness, N. R.; Beton, P. H. Directing Two-Dimensional Molecular Crystallization Using Guest Templates. Chem. Commun. 2008, 2304-2306.

(29) Furukawa, S.; Tahara, K.; De Schryver, F. C.; Van der Auweraer, M.; Tobe, Y.; De Feyter, S. Structural Transformation of a Two-Dimensional Molecular Network in Response to Selective Guest Inclusion. Angew. Chem., Int. Ed. 2007, 46, 2831-2834.

(30) Miao, X.; Xu, L.; Li, Y.; Li, Z.; Zhou, J.; Deng, W. Tuning the Packing Density of Host Molecular Self-Assemblies at the Solid-Liquid Interface Using Guest Molecule. Chem. Commun. 2010, 46, 8830-8832. (31) Lafferentz, L.; Eberhardt, V.; Dri, C.; Africh, C.; Comelli, G.; Esch, F.; Hecht, S.; Grill, L. Controlling On-Surface Polymerization by Hierarchical and Substrate-Directed Growth. Nat. Chem. 2012, 4, 215-220. (32) Schlögl, S.; Heckl, W. M.; Lackinger, M. On-Surface Radical Addition of Triply Iodinated Monomers on Au(111): the Influence of Monomer Size and Thermal

ACS Paragon Plus Environment

Page 22 of 25

Page 23 of 25 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Post-Processing. Surf. Sci. 2012, 606, 999-1004. (33) Peyrot, D.; Silly, F. On-Surface Synthesis of Two-Dimensional Covalent Organic Structures versus Halogen-Bonded Self-Assembly: Competing Formation of Organic Nanoarchitectures. ACS Nano 2016, 10, 5490-5498. (34) Treier, M.; Pignedoli, C. A.; Laino, T.; Rieger, R.; Müllen, K.; Passerone, D.; Fasel, R. Surface-Assisted Cyclodehydrogenation Provides a Synthetic Route Towards Easily Processable and Chemically Tailored Nanographenes. Nat. Chem. 2011, 3, 61-67. (35) Weigelt, S.; Busse, C.; Bombis, C.; Knudsen, M. M.; Gothelf, K. V.; Laegsgaard, E.; Besenbacher, F.; Linderoth, T. R. Surface Synthesis of 2D Branched Polymer Nanostructures. Angew. Chem., Int. Ed. 2008, 47, 4406-4410. (36) Larrea, C. R.; Baddeley, C. J. Fabrication of a High-Quality, Porous, Surface-Confined Covalent Organic Framework on a Reactive Metal Surface. ChemPhysChem 2016, 17, 971-975. (37) Dienstmaier, J. F.; Gigler, A. M.; Goetz, A. J.; Knochel, P.; Bein, T.; Lyapin, A.; Reichlmaier, S.; Heckl, W. M.; Lackinger, M. Synthesis of Well-Ordered COF Monolayers: Surface Growth of Nanocrystalline Precursors versus Direct On-Surface Polycondensation. ACS Nano 2011, 5, 9737-9745. (38) Franc, G.; Gourdon, A. Covalent Networks through On-Surface Chemistry in Ultra-high Vacuum: State-of-the-Art and Recent Developments. Phys. Chem. Chem. Phys. 2011, 13, 14283-14292. (39) Otto, S.; Furlan, R. L. E.; Sanders, J. K. M. Selection and Amplification of Hosts from Dynamic Combinatorial Libraries of Macrocyclic Disulfides. Science 2002, 297, 590-593.

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

(40) Beeren, S. R.; Sanders, J. K. M. Discovery of Linear Receptors for Multiple Dihydrogen Phosphate Ions Using Dynamic Combinatorial Chemistry. J. Am. Chem. Soc. 2011, 133, 3804-3807. (41) West, K.; Bake, K.; Otto, S. Dynamic Combinatorial Libraries of Disulfide Cages in Water. Org. Lett. 2005, 7, 2615-2618. (42) Buryak, A.; Severin, K. Dynamic Combinatorial Libraries of Dye Complexes as Sensors. Angew. Chem., Int. Ed. 2005, 44, 7935-7938. (43) Plas, J.; Waghray, D.; Adisoejoso, J.; Ivasenko, O.; Dehaen, W.; De Feyter, S. Insights into Dynamic Covalent Chemistry at Surfaces. Chem. Commun. 2015, 51, 16338-16341. (44) Ciesielski, A.; El Garah, M.; Haar, S.; Kovaříček, P.; Lehn, J. M.; Samorì, P. Dynamic Covalent Chemistry of Bisimines at the Solid/Liquid Interface Monitored by Scanning Tunnelling Microscopy. Nat. Chem. 2014, 6, 1017-1023. (45) Sun, J.; Yu, Y. X.; Liu, C. H.; Lei, S. B. Surface- and Guest-Promoted Product Selection from a Dynamic Covalent Library: A Scanning Tunneling Microscopic Study. J. Phys. Chem. C 2017, 121, 3437-3444. (46) Lei, S.; Surin, M.; Tahara, K.; Adisoejoso, J.; Lazzaroni, R.; Tobe Y.; De Feyter, S. Programmable Hierarchical Three-Component 2D Assembly at a Liquid-Solid interface: Recognition, Selection and Transformation, Nano Lett. 2008, 8, 2541-2546.

ACS Paragon Plus Environment

Page 24 of 25

Page 25 of 25 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Table of contents

ACS Paragon Plus Environment